Heat Transfer From A Hot Moving Cylinder

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

International Journal of Heat and Mass Transfer 54 (2011) 5527–5539

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Heat transfer from a hot moving cylinder impinged by a planar subcooled water jet
M. Gradeck a,⇑, A. Kouachi a, J.L. Borean b, P. Gardin b, M. Lebouché a
a
LEMTA, Nancy University – CNRS, 2 avenue de la forêt de haye 54504, Vandoeuvre lès Nancy, France
b
AM Research, voie romaine 57283, Maizières lès Metz, France

a r t i c l e i n f o a b s t r a c t

Article history: A hot moving (rotating) cylinder was heated up to 500–600 °C and then was cooled by a planar water jet
Received 10 December 2010 impinging on a line parallel to the symmetry axis. The time dependent wall temperature was measured
Received in revised form 11 July 2011 using embedded thermocouples and the corresponding wall heat fluxes were estimated through an
Accepted 20 July 2011
inverse conduction method. In a recent paper, we showed that cooling rates depend on the subcooled
Available online 16 August 2011
temperature of the jet, the velocity of the jet and the surface-to-jet velocity ratio. Since the initial tem-
perature of the cylinder was higher than the Leidenfrost temperature, we observed all the boiling regimes
Keywords:
from film boiling to nucleate boiling. The objectives of this paper are firstly to describe the current con-
Heat transfer
Inverse solution
ditions which exist in the Run Out Table in hot rolling mills, secondly to review the main experimental
Heat conduction studies dedicated to jet cooling which have led to modelling heat transfer in boiling conditions and finally
Impinging jet to propose new correlations taking into account the velocity of the wall.
Transient boiling Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction be between 200 and 500 °C, depending on steel grade). At the very
beginning of the cooling phase, the temperature of the steel strip is
In the steel industry, cooling on the Run Out Table (ROT) after above the Leidenfrost temperature so film boiling, transition boil-
hot rolling is one of the most difficult process steps in the hot mill ing, critical heat flux (CHF) and nucleate boiling all occur. Control-
strip. The decrease in the temperature needs to be perfectly con- ling the cooling rate and the homogeneity thereof thus remains a
trolled because the mechanical properties of steel alloys are condi- major challenge for manufacturers aiming to produce steels with
tioned by the cooling rate ensured by these jets [1–4]. Generally, desired and homogeneous mechanical properties.
top cooling is carried out using a number of subcooled water jets
which impinge perpendicularly on the hot steel surface while bot-
1.1. Description of the flow on the Run Out Table (ROT)
tom cooling is done using sprays. The water jets are organized in a
set called a header where two jets rows are either aligned or stag-
As previously described, the heat transfer can not be homoge-
gered. Complex flows are thus obtained, as a result of the interac-
neous when using water jets in the cooling system because of flow
tion between the jets and the moving surface. In-depth knowledge
topology and because the water film depth above the hot surface is
on the heat transfer associated to that flow is therefore essential
not constant. This depends on the distance from the impact zone of
including knowledge of the interaction between the jet and the
the jet but also on the ratio between velocity of the jet and the
moving surface, the interaction between the jets and the interac-
velocity of the moving surface. In a recent paper, Gradeck et al.
tion between the ramps.
[5] carried out experimental and numerical studies of the flow
Rates of cooling will vary between 15 and 1000 K/s depending
structure of a single water jet impinging on a moving surface. This
on the required steel mechanical property. Although cooling tech-
work provided a valuable correlation to predict the position of the
nologies based on the so-called laminar water jets have been
hydraulic water jump for operating conditions similar to those in
widely studied in the past, knowledge of these technologies re-
ROT cooling systems. In a more recent paper, the flow pattern of
mains incomplete which means it is difficult to attain optimum
multiple water jet impinging on a moving surface was numerically
production. Obviously the kinetics of cooling depends on the vari-
studied by Cho et al. [6] using the CFD Fluent package. Their com-
ous boiling regimes met during transient cooling (i.e. the metal
putations clearly showed how flow patterns are dependant on the
slab is reheated before rolling, the temperature of the strip after
running conditions (flow rates, velocity of the surface). At low flow
rolling is about 900 °C and after cooling, the temperature should
rates, hydraulic jumps were observed while with increasing values
of the flow rates, the hydraulic jumps disappeared and a pool was
⇑ Corresponding author. observed. Moreover, a fountain effect was found to occur at times
E-mail address: michel.gradeck@ensem.inpl-nancy.fr (M. Gradeck). between two adjacent jets with consequent improvement of the

0017-9310/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2011.07.038
5528 M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539

Nomenclature

a heat transfer coefficient (W m2 K1) r⁄ dimensionless velocity, VVSJ


d nozzle diameter (m) x distance from the jet axis (m)
q00 heat flux (W m2) x⁄ dimensionless distance, dx
Re Reynolds number
Pr Prandtl number Subscripts
DTsub subcooled temperature (K) FB film boiling
DTsat superheat temperature (K) MFB minimum of film boiling
T temperature (°C) TB transition boiling
VJ jet velocity at the impingement (m s1) NB nucleate boiling
Vn jet velocity at the exit of the nozzle (m s1) TC transient conduction
Vs wall velocity (m s1) CHF critical heat flux
k conductivity (W m1 K1) L liquid
l dynamic viscosity (Pa s) V vapour
r surface tension (N m1)

cooling flux. Monde et al. [7] gave an illustration of the interactions (or models) available for impinging jets in these four boiling
of neighbouring jets and the expected increase of the heat flux in regimes.
this area. However in more recent study, Franco [8] did not find
a fountain effect to occur whether the alignment of the jets was 1.3.1. Film boiling
staggered or aligned. The large distance between jets (around 1.3.1.1. Stagnation zone. Zumbrunnen et al. [20,21] and Filipovic
100 mm) in Franco’s [8] experiments is the probable explanation. [22] obtained models for the film boiling regime by solving the
boundary layer equations but as some strong assumptions have
1.2. Heat transfer associated with single impinging jet been assumed to simplify the Navier–Stokes equations and obtain
an analytical solution, these models can be very far from the cur-
Most studies of heat transfer associated with impinging jets rent application. For example neglecting waviness on the va-
have been carried out using a static surface which means that pour–liquid interface as well as Kelvin–Helmotz instabilities lead
the dynamic conditions as mentioned above may have been far to delay the minimum of film boiling (in comparison with
from industrial conditions [9–15]. However, most correlations measurements).
from these studies are used as command and control for the ROT Liu and Wang [23,24] derived an analytical expression of the
even though the influence of surface velocity was not addressed heat flux for the film boiling regime in the stagnation zone
in these studies. Depending on the strip temperature, four different (C = 1.414). To fit the experiments, the constant C of the original
water cooling regimes may be encountered: (i) for high tempera- expression was slightly modified to take into account the interfa-
tures, only film boiling; (ii) transition boiling if the surface temper- cial waviness of the vapor layer (C = 2):
ature of the strip is lower than the Leidenfrost temperature; (iii) .
nucleate boiling regime if the surface temperature is lower than Q 00FB ¼ CRe0;5 0166
J Pr J ðkL kV DT sub DT sat Þ0;5 d; ð1Þ
the CHF temperature and finally; (iv) forced convection
(Fig. 1(a)). A particular shape of boiling curve may be observed in where the thermal properties of water were evaluated at the film
the transition regime in the case of an impinging jet on a hot plate temperature of water and vapor properties evaluated at the film
(Fig. 1(b)). After CHF, a first minimum of flux was observed after temperature of vapour.
which the heat flux increased again and reached a high value. This Ochi et al. [10] proposed the following correlation
‘‘shoulder of flux’’ was found to falls down abruptly to film boiling q00FB ¼ 3:18x105 ð1 þ 0:383DT sub ÞðV J =dÞ0:828 ð2Þ
regime for very high superheats. Miyasaka et al. [15], Ishigai et al.
[9], Ochi et al. [10] and Hall [11], Robidou et al. [13] and Gradeck et A similar expression was found by Ishigai et al. [9] for a planar jet:
al. [16] have reported the existence of a ‘‘shoulder of flux’’ beneath
the jet (transition boiling regime) in the case of static surface. But q00FB ¼ 5:4x104 ð1 þ 0:527DT sub ÞV 0:607
J : ð3Þ
Gradeck et al. [16] showed that heat transfer is radically modified 1
For all of these correlations V is expressed in m s , d is in mm and
when the impingement surface is moving. They found that the
DT in K.
‘‘shoulder of flux’’ beneath the jet axis collapsed and that the local
Robidou [25] used an innovative experimental device to mea-
boiling curves had the same shape upstream, downstream and be-
sure boiling curves under steady state conditions and proposed
neath the centreline of the jet.
the following correlation at the stagnation zone:

1.3. ROT heat transfer model q00FB ¼ 5:38x104 ð5:5 þ DT sub ÞV 0:6 ð4Þ
J

Given that cooling rates (i.e., heat transfer at the wall) are af-
fected by the temperature of the strip, dedicated correlation needs 1.3.1.2. Parallel flow zone. Few models exist in the parallel flow
be applied to the off-line or on-line models used in ROT in order to zone (x⁄ > 2). Filipovic [22] developed a model for this regime
correctly predict the coiling temperature of the strip. Since the heat based on the boundary layers equations and measurements
transfer model used in the ROT control is usually simple, consisting achieved on a parallel jet. Hatta et al. [26] proposed the following
of a simple polynomial relationship [17–19], improvements should correlation fitting the experimental data of Kokado et al. [27] ob-
be done to consider four heat transfer regimes (Fig. 1) and two sig- tained for a transient cooling:
nificant temperatures (Leidenfrost and CHF temperatures). In the
following sub-sections, we give an overview of the correlations a ¼ 200ð2420  21:7T L ÞDT 0:8
sat : ð5Þ
M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539 5529

(a)

Forced Nucleate Transition Filmboiling


convection boiling boiling

Heat flux (W/m²)

Critical heatflux

Leiden frost point

ΔTsat = Tp-Tsat (K) (K)

(b)

Forced Nucleate Transition Film boiling


convection boiling boiling
Heat flux (W/m²)

Critical heat flux

Leiden frost point

ΔTsat = Tp-Tsat (K)

Fig. 1. Shape of the boiling curve in impinging jet experiments: (a) parallel flow zone of the jet, (b) impinging zone of the jet.

This correlation underestimated the heat transfer coefficient in T MFB ¼ a  bT L ; ð8aÞ


comparison with the values measured by Robidou [25] at the min-
imum of film boiling (MFB). with a = 1100 °C, b = 8.5 for Kokado’s experiment.
The previous relationship must be used according to the critical
1.3.2. Temperature for the minimum film boiling (MFB) temperature, Tc (which was 68 °C in the case of Kokado’s experi-
Several results can be found concerning the temperature for the ments): (i) If TL < Tc, a wetting regime occurs; (ii) if TL > Tc and
MFB (i.e. TMFB) in the stagnation zone. Liu [24] gave an expression TW < TMFB, a partial wetting regime occurs; (iii) if TL > Tc and
for the temperature of minimum film boiling, TMFB. TW > TMFB, film boiling can occur. Following on from these results,
  under the jet and for water temperature below 40 °C, the wall is al-
DT MFB ¼ 14DT sub Pr0;33
J ðkL =kV Þ lL =lV ð6Þ ways in a wetted regime.
Hatta and Osakabe [28] used the water temperature whereas
And the associated heat flux:
Robidou [25] used the subcooling:
 0;5
q00FB;h ¼ 0:5DT MFB kV Re0;5 0166
J Pr J lL = lV =d; ð7Þ
T MFB ¼ 326 þ 17:6DT 0:8
sub ð8bÞ
where the thermal properties of water were evaluated at the film
temperature of water and vapor properties evaluated at the film A change of TMFB according to the hydrodynamics of the jet-wall
temperature of vapor. Following Liu [24], TMFB depends only on interaction needs to be taken into account in this expression.
water properties and subcooling and hydrodynamics does not have In a recent paper, Karwa et al. [29] have also developed a sim-
any effect (see expression 6), Kokado et al. [27]: plified two-phase analytical model for predicting of the TMFB in
5530 M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539

the case of a planar or a circular liquid jet impinging on a flat hot Miyasaka et al. [15] found the following relationship to estimate
wall. They provided the following equation: CHF for a planar jet impinging a static plate:

DT MFB ¼ 0:7DT sub Pr0;5 0;5


L ReL ðkL =kV ÞðdV =wÞ for a planar jet; ð9aÞ q0CHF ¼ 0; 16hlv qv ð1 þ 0; 86V 0;38
n Þ
 0;8  1;13 ! 
0;5
ql cpl DT sub rgðql  qv Þ 0;25
DT MFB ¼ 1:6DT sub Pr 0;5
L ðkL =kV ÞðlL =lV Þ for a circular jet; ð9bÞ  1 þ 0112
qv h lv q2v
:

where the properties of vapour are taken at the bulk temperature, w ð12bÞ
is the width of the jet just before impingement and dV, the vapour
thickness. Although the hydrodynamic parameters used are not too far from
the requirements of the ROT, several points prevent the extension
1.3.3. Transition boiling of these correlations to the ROT including:
Transition boiling is an important regime because it is at the
origin of control issues. When the strip temperature decreases, ? The surface is fixed.
the extracted heat flux increases and the temperature discrepan- ? Experiments are dependent on the size of heater, which is a
cies are amplified because of high discrepancies on local heat flux. parameter of the correlations.
A definition of the transition boiling was provided by Berenson
[30], Kalinin et al. [31] and Pan et al. [32] namely a combination 1.3.5. Nucleate boiling
of unstable film boiling and unstable nucleate boiling alternately Wolf [12], Wolf et al. [37] proposed correlation (9) for nucleate
existing at any given location on a heating surface. The extracted boiling:
heat flux in the transition boiling was provided by Kalinin et al.
q00NB ¼ 63:7DT 2:95
sat : ð13Þ
[31]:
q00TC tTC þ q00NB t NB þ q00FB t FB Following Wolf’s finding [12], the previous correlation can also be
q00TB ¼ ð10Þ applied in a parallel flow configuration:
tTC þ t NB þ tFB
With TB, transition boiling; TC, transient conduction; NB, saturated q0NB ¼ aDT bsat : ð14Þ
nucleate boiling; FB, film boiling;t, fraction of time, contact time of This correlation means that heat transfer does not depend on
each regime. hydrodynamics in this regime.
For Nagai and Nishio [33], Ohtake and Koizumi [34], transition Filipovic et al. [38,39] proposed the following correlation (for a
boiling is a stationary boiling regime and the heat flux is expressed static surface) taking into account the velocity of the jet, subcool-
as: ing, the diameter of the nozzle and superheat. However this corre-
q00TB ¼ Cq00NB þ ð1  CÞq00FB ð11Þ lation also overestimates the heat flux in the case of nucleate
boiling for moving system:
In fact both approaches lead to the last equation with the weighting
0:608
parameter having a different meaning when transient conduction is q0NB ¼ C ðDT sub ÞV 0:5
j d DT 0:14
sat ; ð15Þ
assumed as negligible. In the case of time-dependant models, the 7
where C(DTsub = 80 °C) = 1.3  10 and C(DTsub = 65 °C) = 0.81 
weighting C is based on contact time whereas it is based on the li-
107; Vj in m/s, d in mm.
quid–solid contact fraction area in stationary models. These ap-
proaches are very interesting because expressions can be adapted
1.4. Concluding remarks on our literature review
to the case of the ROT using experiments. The current parameters
used are obtained for nuclear reactor applications by using droplets,
The real process is extremely complicated because of the multi-
pool boiling experiments, vertical tubes, saturated fluid, tempera-
ple interactions involved namely the interaction of the jet with the
tures below 300 °C, with surfaces (copper, sapphire, etc.) whose
band, of a jet with another jet or of a header with another header.
thermal properties are very different from those of steel. These can-
This means it is impossible to cover all these situations at the same
not be used for the Run-Out Table without prior adaptations. More-
time in the context of a simple laboratory experiment. Hence pre-
over the moving effect of the surface is not taken into account at all
vious studies have always been carried out using simple configura-
although Gradeck et al. [16] found this significantly modifies the
tions such as multiple jet impinging a cold plate, a single jet
boiling curve and transition boiling regime.
impinging a moving cold plate or a single jet impinging one plate
whose temperature was higher than the Leidenfrost temperature.
1.3.4. Critical heat flux, CHF
Previous theoretical approaches also used a simple situation as
Monde et al. [7] and his team (Mitsutake et al. [35], Hammad
their starting point, for example an axisymmetric jet impinging a
et al. [36]) have carried out a lot of experiments in the field of
static cold surface or planar jet impinging a moving plate (cold
impinging jets and they have provided a correlation to estimate
or hot). For our study, we opted to experiment with a new situa-
the CHF:
8 pffiffiffiffiffiffiffiffiffiffiffi tion, namely a planar jet impinging a hot moving surface (i.e. a
>
> q00
¼
1þ 1þ4CJa rotating cylinder). The Table 1 gives an overview of the running
> qCHF
> 2
;
>
> 0
conditions for the different experimental works which have been
>
> q CpDT
< Ja ¼ qVL hLVsub ; previously mentioned.
 0:645  0:343 ; ð12aÞ
>
> q0 ¼ 0:221 qqL 2r
ð1 þ D=dÞ0:364 ;
>
> qL V J ðDdÞ
>
>
V 2. Experimental set-up and data reduction
>
> 0:95ðd=DÞð1þD=dÞ0:364
:C ¼
ðq =q Þ0:43 ð2r=q V ðDdÞÞ0:343
L V L J
The experimental device was designed to allow local and instan-
where D was the size of the heater and thermal properties were taneous determination of the heat transfer coefficient between a
evaluated at the saturated temperature. This correlation is applica- water jet impinging a moving surface. This estimation was quite
ble for stationary applications, static surface and to a qqL of 5.3–1603, accurate because we used temperature measurements of the wall
V
the inverse Weber number (second term in brackets) of 2  107 to surface and calculating the cooling heat flux by means of an inverse
3
10 and D/d of 5–30. conduction method [40]. This device is illustrated in Fig. 2.
M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539 5531

Table 1
Experimental conditions of the cited works.

Authors Experiment Liquid Jet Surface Running conditions Regimes


2
Ishigai et al. [9] Transient Water Free (20 or 80)  12 mm 1 6 VJ 6 3.2 m/s Nucleate boiling
Planar Stainless steel 5 6 DTsub 6 55 K Transition boiling
50  6.2 mm2 Film boiling
Ochi et al. [10] Transient Water Free 210  50 mm2 2 6 VJ 6 7 m/s Nucleate boiling
Circular Stainless steel 5 6 DTsubB 6 80 K Transition boiling
5, 10 and 20 mm Film boiling
Kokado et al. [27] Transient Water Free 2000  10 mm2 1 6 QV 6 7 l/min Wetting
Circular Stainless steel TL = 20 °C Non wetting
Ø10 mm
Mitsutake et al. [35] Transient Water Free Width 4 mm 5 6 VJ 6 60 m/s CHF
Circular Thickness 0.1 mm 0.1 6 P 6 1.3 MPa
Ø2 mm Stainless steel TL = 20 °C
Ni
Wolf [12] Stationary Water Free 119  35.7 mm2 10 6 x 6 90 mm Forced convection
Imposed heat flux Planar Alloy Ni–Cr–W–Mo 2 6 VJ 6 5 m/s Onset of nucleate boiling
10  105 mm2 DTsub = 50 K
0.25 6 q00 6 6.34 MW/m2 Nucleate boiling
Miyasaka et al. [15] Stationary Water Free 4  8 mm2 1.1 6 VJ 6 15.3 m/s Forced convection
Planar Pt TL = 15 °C Nucleate boiling
10  32 mm2 CHF
Miyasaka et al. [15] Stationary Water Free Ø1.5 mm 1.1 6 VJ 6 15.3 m/s Nucleate boiling
Planar Pt TL = 15 °C CHF
10  32 mm2 Transition boiling
Robidou et al. [13] Stationary Water Free 10  104 mm2 0.46 6 Vn 6 0.9 m/s All regimes
Controlled temperature Planar Ni on Cu 5 6 DTsub 6 17 K
1  9 mm2
Kouachi [41], Transient Water Free Rotating cylinder 0.32 6 Vn 6 0.69 m/s All regimes
Gradeck et al. [16]
Planar Ni 0.5 6 r⁄ 6 1.25 m/s
4  180 mm2 Ø175 mm 18 6 DTsub 6 50 K

Inspection 5 Planar Jet


Glass
6 6
8 8
ω

7
24T 9 10

11

T: temperature measurement
Q: flow rate measurement
ω: velocity measurement (rpm)
P: power measurement

T
Q

Fig. 2b. Details of the experimental set-up [41].


2 3 4
1
(flow rate and water temperature). The results obtained in this
Fig. 2a. Schematic of the experimental set-up [41]. study of heat transfer beneath the jet and close to the impact zone
(in these zones, the effect of the curvature is negligible) can used
2.1. Experimental set-up for ROT characterisation. However, this device cannot be used to
simulate the effect of the accumulation of water because in the real
The cooling of the steel strip in the ROT can be approximated process a certain amount of water can be transported by the steel
using a rotating hot cylinder which respects the main process strip. Of course, water accumulation is also a parameter that
parameters of the cooling to some extent namely the speed of should be considered in future works because heat transfer may
the moving surface, its temperature and the jet characteristics be different for a free jet or a submerged jet.
5532 M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539

Sensor signal
T 2Τ 3T time =

q’’ (W/m²)
Temperature 2π or x*=0 2π 2π position

time

ΔTSAT
Heat flux

time

Fig. 3. Building of the boiling curves.

The set-up is schematically outlined in Fig. 2a and photos 2.2. Data reduction
show details of the device in Fig. 2b. The boiling vessel contained
the jet (5) and the hot rotating cylinder (7). The external When the cylinder reached the equilibrium temperature, the
diameter of the outer cylinder (which is cooled by the jet) was heating was stopped and the jet cooling began at t = 0 (opening
equal to 175 mm and its internal diameter equals 100 mm. The of the sluicegate). During cooling, the heat flux was estimated
length of this cylinder was 200 mm. Heating was provided by through inversion of the temperature measurements using an In-
three electrical wires which were wound on an inner cylinder verse Heat Conduction Model (IHCM); see Volle et al. [40]. This
(Øout ¼ 100 mm;Øin ¼ 85 mm) which itself was adjusted with model was based on an analytical solution of the 2D transient heat
the outer cylinder. Heating with a constant heat flux (P) was thus equation. As axial conduction was neglected, we therefore worked
applied on the internal radius until thermal equilibrium was with a 2D problem. The heat equation can be solved using a cylin-
reached (uniform wall temperature required for the test of about drical reference with knowledge of the initial temperature field
600 °C on the external radius). Each resistance used can dissipate (i.e. an equilibrium temperature field obtained with a constant
2000 W (10). The wall temperature was almost uniform because heat flux P at an internal radius with losses due to convection
heat losses (radiation and convection) were almost uniform (i.e. and radiation at the external surface and conduction at each end
the cylinder rotates during the heating phase). Two cooled stain- of cylinder). The use of spatial integral transform, see Maillet
less steel flasks (6) were set out on both sides to prevent the re- et al. [42], namely the Fourier transform (noted), led us to obtain
volving contacts, ball bearings and sealing rings from excessively a linear relation (in the Fourier space) between the kth harmonic of
high temperatures. This was also used to center the cylinder and inside temperature1 and the kth harmonic of heat flux lost on the
thus avoid the unbalances. The system was allowed to dilate cooled wall:
without generating additional constraints inside the set-up. Axial
losses (i.e. axial conduction) are impossible to avoid but as the
~ kq
T~ k ¼ X ~k ; ð16Þ
cylinder was long enough and thermocouples placed in the mid- ~ k , the sensitivity matrix linking variations of temperature re-
with X
dle of it, we only considered a 2D conduction problem to solve
sponses to variations of wall heat flux input. Once T~ k was estimated
the inverse problem related to the estimation of heat flux [40].
through experimental measurements, q ~k could be calculated thanks
24 thermocouples (N type) were inserted near the external sur-
to an inversion of the previous equation. Because of the standard
face in grooves and a coating was deposited. The electrical output
noise deviation of experimental temperatures, an ordinary least
was provided to resistances by means of revolving contacts (8a).
square operator was used to obtain the most accurate estimation
The test fluid (water) was first heated at a subcooled temperature
of the cooling heat flux in Fourier domain:
in a vessel (1) and circulated in a primary circuit. The flow rate
was adjusted using a pump (2) and controlled using an electro-  1
~k ¼ X
q ~
~T X ~ T T~ k :
X ð17Þ
magnetic flowmeter (3). When all the initial parameters were ad- k k k
justed, the electromagnetic sluice gate (4) was opened and the
cooling of the cylinder could start. The jet had an outlet cross A Fourier inversion of the previous estimated signal led to the cool-
sectional area of 180  4 mm2 and the water jet outlet was lo- ing flux identification as a function of time. Obviously the inverse
cated at a distance H from the hot cylinder and impinged the problem is an ill-posed problem and it was thus necessary to use
higher line of the cylinder. This distance could be adjusted. The some regularization techniques. In our study, the probes were lo-
pre-amplified signals of the thermocouples as well as their angu- cated close to the wall (500 lm depth) and we used a Beck future
lar positions (i.e. a position sensor was available on the drive- time regularization method to obtain a better estimation. The accu-
shaft) were then also transmitted to a computer (11) via racy of the method was estimated using noisy data (i.e. tempera-
revolving contacts (8b) to be recorded. This meant we could con- ture) which was obtained either using the analytical solution of
struct the time evolution of temperature at each point on the the heat equation or numerical simulations. Any errors would
surface of the cylinder. More details on this device can be found
in [41]. 1
The thermocouples were inserted at 500 lm depth.
M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539 5533

depend on the choice of hyperparameters or regularization param- heat flux (Fig. 4c), first minimum of film boiling and shoulder of
eters (number of future times, number of harmonics) and of course flux (Fig. 4a).
on the noise level [43,44].
As we could know the position of the thermocouple as a func-
tion of time, it was then possible to link the time evolution of 3. Heat transfer analysis and correlations
the temperature at a given point to the time-evolution of heat flux
at the same point and thus to build the boiling curve corresponding 3.1. Heat transfer analysis
to a curvilinear abscissa x⁄ (Fig. 3). An example of the data obtained
(local boiling curves) is given in the Fig. 4. Thus we were able to As described in a previous paper [16], heat transfer does not
identify the different heat transfer regimes namely forced convec- evolve in the same way as in a static case and its value is clearly
tion, onset of nucleate boiling, nucleate boiling (Fig. 4b), critical modified when the impingement surface is moving. As a

6
x 10
3
x* = 6.9
x* = 9.15
2.5 x* = 11.45
x* = 13.75
x* = 16
2
Heat flux (W/m²)

1.5

0.5

0
0 50 100 150 200 250 300 350 400 450
DTsat (K)
(a)
5 6
x 10 x 10
3.5
x* = 6.9 x* = 6.9
16
x* = 9.15 CHF at MHF x* = 9.15
x* = 11.45 x* = 11.45
14
Heat flux (W/m²)

3
Heat flux (W/m²)

x* = 13.75 x* = 13.75
nucleate boiling x* = 16 x* = 16
12

10 ONB 2.5

8
2
6 Forced convection

4
1.5
2

0 1
10 20 30 40 50 60 70 40 50 60 70 80 90 100 110 120 130
DTsat (K) DTsat (K)

(b) (c)

legend
Fig. 4. Local boiling curves from [16], DTsub = 34 K; uj = 1,06 m/s  u⁄ = 1,25.
5534 M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539

because the jet was spreading over a stable vapour film. Neverthe-
less, in a recent published paper [16], we especially shown a col-
q’’ (W/m²)

Increasing the surface velocity


lapse of the ‘‘shoulder of flux’’ beneath the jet axis and a
displacement of the maximum of heat transfer downstream. An-
other point is that the local boiling curves have slightly the same
shape upstream, downstream and beneath the point of maximum
heat transfer (i.e. maximum heat transfer was not always at the
centreline of the jet; see Gradeck et al. [16]). The shape of the boil-
ing curve looked rather like those beneath the impingement
(Fig. 5). Seiler-Marie et al. [45] have modelled the transition, from
transition boiling to the shoulder of flux and this is based on the
assumption of the existence of periodic bubble oscillations con-
First minimum Leidenfrost temperature
trolled by Rayleigh–Taylor instabilities namely that vapour spots
would be fragmented by the jet hydrodynamic forces. In fact, for
ΔTSAT (K) moving surfaces, rewetting the wall and thus stability of the va-
pour spots or film should be also controlled by Kelvin–Helmholtz
Fig. 5. Shape of the local boiling curves with a moving surface.
instabilities.
Another point to be recalled is related to the position of the
consequence, the maximum of CHF is lower than in the static case cooling zone (i.e. the zone corresponding to a temperature de-
[25]. But the most obvious effect was a kind of auto-similarity of crease): at the beginning of cooling, the cooling zone begins mainly
the local boiling curves throughout the cooling zone. In fact, defin- beneath the jet and develops only downstream [16]. As a conse-
ing a local boiling curve at x⁄ for a moving surface integrates all the quence, the maximum of heat transfer was not located beneath
boiling phenomena over the surface throughout the cooling zone the jet and it moves during the cooling. An asymmetry of the heat
upstream. Bubbles or vapour spot moved with the surface (for transfer was also observed upstream and downstream from the
nucleate boiling and transition boiling regimes) and in the film maximum of heat transfer (Fig. 4). This was a consequence of the
boiling regime, the stability of the vapour layer is clearly hard to jet-wall interaction, especially when a counter current appeared
achieve which is why a kind of ‘‘shoulder of flux’’ was observed. (upstream) and induced rapid evolution of the velocity flow field.
For a static surface, the shoulder of flux was observed at the The data bank obtained by Kouachi [41] was analyzed and heat
impingement of the jet (x⁄ = 0) but collapsed in the flow zone transfer correlations for nucleate boiling and maximum heat

4 4
Vj = 1,06 m/s; r* = 0,5; ΔTsub = 18 K x=0 Vj = 1,06 m/s; r* = 1; DTsub = 34 K
x= 4,5 Vj = 1,06 m/s; r* = 1; DTsub = 18 K
3.5 3.5
x = 9,15

3 3
Heat flux (MW/m²)
Heat flux (MW/m2)

2.5 2.5

2 2

1.5 1.5

1 1

0.5 0.5

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
ΔTsat (K) ΔΤsat (Κ)

(a) local boiling curves (b) Influence of subcooling, x/d = 9.15


4 4
Vj = 1,20m/s; Vs = 1,14 m/s; DTsub = 50 K
Vj = 1,06 m/s; ΔTsub = 50 K
r* = 0,5
3.5 3.5
Vj = 1,04m/s; Vs = 1,14 m/s; DTsub = 50 K r* = 1
r* = 1,25
3 3
Heat flux (Mw/m²)

Heat flux (MW/m²)

2.5 2.5

2 2

1.5 1.5

1 1

0.5 0.5

0 0
0 20 40 60 80 100 0 20 40 60 80 100
ΔTsat (K) ΔTsat (K)

(c) Influence of jet velocity, x/d = 9.15 (d) Influence of speed surface, x/d = 8
Fig. 6. Local evolution of boiling curves in nucleate boiling regime.
M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539 5535

2.5

+20%

Calculated heat flux (MW/m²)


2 -20%

1.5

0.5

0
0 0.5 1 1.5 2 2.5
Experimental data (MW/m²)

Fig. 7. Comparison between experimental data and correlation (18).

6
4
Vj = 1,06 m/s; DTsub = 50 K r* = 0,5 Vj = 1,06 m/s; r* = 1,25
r* = 0,75 DTsub = 50 K
r* = 1 3.5 DTsub = 34 K
5 r* = 1,25
DTsub =18 K
3
4
CHF (MW/m²)

2.5
CHF (MW/m²)

3 2

1.5
2

Vs 1
1
0.5

0 0
0 2 4 6 8 10 12 14 16 18 0 2 4 6 8 10 12 14 16 18
x/d x/d

(a) influence of the surface velocity (b) influence of the subcooling


6 6
Vs = 1,14 m/s; DTsub = 50 K Vj = 0,69 m/s; DTsub = 50 K
H = 50 mm
Vj = 1,04 m/s
5 H = 10 mm
5 Vj = 1,20 m/s

4 4
CHF (MW/m²)

CHF (MW/m²)

3 3

2 2

1 1

0 0
0 2 4 6 8 10 12 14 16 18 0 2 4 6 8 10 12 14 16 18
x/d x/d

(c) influence of the jet velocity (d) influence of the nozzle distance
Fig. 8. Local evolutions of the CHF.

transfer were proposed taking into account the main hydrody- 3.2. Heat transfer associated with moving surface
namic parameters, the fluid properties, the fluid subcooling and
also the dissymmetry of the heat transfer between upstream flow 3.2.1. Nucleate boiling regime
and downstream flow. Further experiments would be required to The heat flux at the onset of boiling was independent of
study other regimes like transition boiling or film boiling. curvilinear position but slightly increased with subcooling. In the
5536 M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539

15

14
+12%
13

12

calculated x/d
11

10 -12%

5
5 7 9 11 13 15
experimental x/d

Fig. 9. Comparison between data and calculated xCHF;max using (21).

6
+20%
5
calculated CHF (MW/m²)

4 -20%

0
0 1 2 3 4 5 6
experimental CHF (MW/m²)

Fig. 10a. Comparison between data and calculated CHF using (22), upstream.

+20%
5
calculated CHF (MW/m²)

4 -20%

0
0 1 2 3 4 5 6
experimental CHF (MW/m²)

Fig. 10b. Comparison between data and calculated CHF using (22), downstream.
M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539 5537

12
Wolf [38]
Gradeck et al.
10

Heat flux (MW/m²)


8

0
20 30 40 50 60 70
ΔTsat (K)

Fig. 11a. Comparison of the correlations (12) and (17) for nucleate boiling.

25
Miyasaka et al. [15]
Gradeck et al.
20
Heat flux (MW/m²)

15

10

0
0 20 40 60 80 100
ΔTsub (K)

Fig. 11b. Comparison of the correlations (12b) and (22) for CHF, x⁄ = 0, r⁄ = 0.

nucleate boiling regime, the curvilinear position and subcooling 3.2.2. CHF
was found to have little influence on the cooling flux (Fig. 6a and The CHF can be observed only for x⁄ > 0 [16]. No cooling was
b). measured for negative values of x⁄. Fig. 8a and b show the
Fig. 6c and d, respectively show the influence of jet velocity and evolution of local CHF for various conditions of speed and
the influence of wall velocity on the boiling curves. The cooling flux subcooling.
at the onset of boiling varied slightly with the wall velocity but in- The CHF was not found to be uniform throughout the cooling
creased quicker with the speed of the jet. The cooling flux in the zone and thus there was a maximum that decreased when the
fully developed nucleate boiling regime seemed independent of wall speed was increased (Fig. 8a). Moreover, the position of that
jet velocity and wall velocity. However, it should be noted that maximum increased when the wall speed was increased and
near the CHF (DTsat around 80–100 K), the dependence of the cool- moved towards high x⁄ values. Another point to be noted was
ing flux on the speed surface became significant, Fig. 6d. that the CHF was also controlled by the water subcooling. In
This means that in the fully developed nucleate boiling regime Fig. 8b, we plotted its spatial evolution for three values of
heat transfer was found not to depend on hydrodynamics which subcooling.
was expected. This means our data could be fitted with a power Fig. 8c and d show the evolution of CHF for different jet veloci-
law of DTsat. The correlation that best fits our data in the nucleate ties (i.e., increasing the distance between nozzle and wall was
boiling regime is: equivalent to increase the jet velocity).
We can define two kinds of correlations as follows:
q0FNB ¼ 2; 46DT 1;64
sat ð18Þ
 
where q0FNB (kW/m2) and 30 < DTsat < 60 K. xCHF;max ¼ f1 V j ; V s ; DT sub ð19Þ
A comparison of the data is shown in the Fig. 7. The closer we  
q0CHF ¼ f2 V j ; V s ; DT sub ; x 
ð20Þ
got to the CHF, the less data fitted with the Eq. (17).
5538 M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539

A dimensional analysis gave the following relationship: up and used to obtain the time-dependent surface temperature
 0794  0209  0247 !0134 and extracted heat flux. The findings of this experimental study
xCHF;max Vs H qv hlv could be used for control cooling systems in ROT. The main find-
¼ 1; 0015We0202 Ja0579 ;
d Vn d ql V 2n ings of this study are as follows:
ð21Þ
c DT (1) The cooling rates were significantly different for moving sur-
with We ¼ 0036  0; 17; Ja ¼ pl h sub ¼ 0033  0088; V s =V n ¼ 1; 36  3;
lv faces as compared with static surface. We observed the same
56; H=d ¼ 2; 5  12; 5; qv =ql ¼ 8; 4  105  33  105 ; hlv =V 2n ¼ 5
6 6 spatial pattern with film boiling, shoulder of flux, transition
10  23  10 .
boiling, CHF and finally nucleate boiling.
Fig. 9 compares the positions of the maximum CHF determined
(2) The main difference was observed in the shoulder of flux
experimentally with those calculated using the previous Eq. (19).
region where the moving of the surface induced a collapse
This shows that the equation correlates the data with a maximum
of the heat flux.
error of ±12%.
(3) Correlations for nucleate boiling and CHF are available for
As we knew the position of the maximum CHF, we could esti-
wide range of operating conditions. Surface velocity is a very
mate the value of CHF upstream and downstream of this position.
sensitive parameter for CHF.
The correlation given by Miyasaka et al. [15] could be corrected by
(4) Finally, we conclude that further experiments are needed to
a factor taking into account the effect of the speed surface (r⁄) as
explain the collapse of the shoulder of flux and to predict the
well as the decrease of CHF from its maximum value:
  rewetting (TMFB) of the surface for high superheats.
x < xCHF;max : q0CHF ¼ 0; 186hlv qv 1 þ 0; 86V 0;38
n
 0;56 !  
q rg ql  qv 0;25 References
 1 þ 0; 328 l Ja
qv 2 qv
 [1] S. Serajzadeh, Prediction of temperature distribution and phase transformation
x 0317 on the run-out table in the process of hot strip rolling, Appl. Math. Model. 27
 1þ ð1 þ r Þ1926 ð22aÞ (2003) 861–875.
d
[2] Y.K. Sun, D. Wu, Effect of ultra-fast cooling on microstructure of large section
  bars of bearing steel, J. Iron. Steel Res. Int. 16 (5) (2009) 61–65. 80.
x > xCHF;max : q0CHF ¼ 0; 186hlv qv 1 þ 0; 86V 0;38
n [3] R.W.K. Honeycombe, Steels: Microstructure and Properties, Edward Arnold,
  !  0;25 London, 1981.
0;56
ql rg ql  qv [4] N. Hatta, J. Kokado, H. Takuda, J. Harada, K. Hiraku, Modelling for cooling
 1 þ 0; 328 Ja process of a hot plate by a laminar water bar, Steel Res. 55 (1984) 143–148.
qv q2v [5] M. Gradeck, A. Kouachi, A. Dani, D. Arnoult, J.L. Borean, Experimental and
 x 0;0486 numerical study of the hydraulic jump of an impinging jet on a moving surface,
 1þ ð1 þ r  Þ0;83 ð22bÞ Exp. Therm. Fluid Sci. 30 (2006) 193–201.
d [6] M.J. Cho, B.G. Thomas, P.J. Lee, Three-dimensional numerical study of

with Vn = 0,32  0,69 m/s; DTsub = 18  50 K et r = 0,5  1,25. impinging water jets in runout table cooling processes, Metall. Mater. Trans.
B 39 (2008) 4.
The correlations correspond to our experimental data with a [7] M. Monde, H. Kusuda, H. Uehara, Burnout heat flux in saturated forced
maximum deviation of ±20% (Figs. 10a and 10b). convection boiling with two or more impinging jets, Trans. JSME 46 (1980)
1834–1843.
[8] G. Franco, Boiling heat transfer during cooling of a hot moving steel plate by
3.3. Comparisons with existing correlations multiple top jets, Ph.D., University of British Columbia, Vancouver, December
2008.
It is clear that heat fluxes were strongly modified in all the boil- [9] S. Ishigai, A. Nakanishi, T. Ochi, Boiling heat transfer for a plane water jet
impinging on a hot surface, in: 6th International Heat Transfer Conference, vol.
ing regimes when the wall was moving. If we compare our results 1, FB 30, 1978, pp. 445–450.
with the correlations given in Section 1, the correlations used so far [10] T. Ochi, S. Nakanishi, M. Kaji, S. Ishigai, Multi-phase and Heat Transfer III. Part
for static surfaces lead to an overestimation of heat transfers for a A: Fundamentals – Cooling of a Hot Plate with an Impinging Circular Water Jet,
Elsevier Science Publishers B.V., Amsterdam, 1984. pp. 671–681.
moving surface. Figs. 11a and 11b show our comparisons of the [11] E.D. Hall, F.P. Incropera, R. Viskanta, Jet impingement boiling from a circular
proposed correlations for nucleate boiling (Eq. (18)) and CHF (Eq free-surface jet during quenching: part 1. Single-phase jet, J. Heat Transfer 123
22) with the findings of Wolf et al. [37] (Eq. (13)) and Miyasaka (2001) 901–910.
[12] D.H. Wolf, Turbulent development in a free surface jet and impingement
et al. [15] (Eq. (12b)). The question was why heat transfer is over-
boiling heat transfer, Ph.D., Purdue University, 1993.
estimated by previous correlations? It is surprising that heat trans- [13] H. Robidou, H. Auracher, P. Gardin, M. Lebouché, Controlled cooling of a hot
fer is widely modified for nucleate boiling because as we showed in plate with a water jet, Exp. Therm. Fluid Sci. 26 (2-4) (2002) 123–129.
[14] Aloke Kumar Mozumder, Masanori Monde, Peter Lloyd Woodfield, Md.
Section 1, most of the correlations used for nucleate boiling do not
Ashraful Islam, Maximum heat flux in relation to quenching of a high
usually take any hydrodynamics effects into account. To our temperature surface with liquid jet impingement, Int. J. Heat Mass Transfer 49
knowledge, only the equation proposed by Filipovic et al. [38,39] (17–18) (2006) 2877–2888.
(correlation for a static surface) takes into account the velocity of [15] Y. Miyasaka, S. Inada, Y. Owase, Critical heat flux and subcooled nucleate
boiling in transient region between a two-dimensional water jet and a heated
the jet, the subcooling, the diameter of the nozzle and superheat surface, J. Chem. Eng. Jpn. 13 (1980) 29–35.
but this also leads to the heat flux in the case of nucleate boiling [16] M. Gradeck, A. Kouachi, M. Lebouché, F. Volle, D. Maillet, J.L. Borean, Boiling
for moving system being overestimated. curves in relation to quenching of a high temperature moving surface with
liquid jet impingement, Int. J. Heat Mass Transfer 52 (5–6) (2009) 1094–1104.
CHF was also overestimated by the correlation given by Mya- [17] Yi. Zheng, Li. Shaoyuan, Xiaobo Wang, Distributed model predictive control for
saka et al. [15]. In that particular case, this is less surprising be- plant-wide hot-rolled strip laminar cooling process, J. Process. Control 19
cause as Critical Heat Flux occurs, the amount of vapour which is (2009) 1427–1437.
[18] Ananya Mukhopadhyay, Sudipta Sikdar, Implementation of an on-line run-out
produced at the wall may be carried by the moving wall which table model in a hot strip mill, J. Mater. Process. Technol. 169 (2005) 164–172.
leads to heat transfer being limited. [19] H.B. Xie, Z.Y. Jiang, X.H. Liu, G.D. Wang, A.K. Tieu, Prediction of coiling
temperature on run-out table of hot strip mill using data mining, J. Mater.
Process. Technol. 177 (2006) 121–125.
4. Concluding remarks [20] D.A. Zumbrunnen, F.P. Incropera, R. Viskanta, Method and apparatus for
measuring heat transfer distributions on moving and stationary plates cooled
This paper presents new results regarding the water cooling of a by a planar liquid jet, Exp. Therm. Fluid Sci. 3 (1990) 202–213.
[21] D.A. Zumbrunnen, F.P. Incropera, R. Viskanta, The effect of surface motion on
hot moving strip which was initially at a temperature around forced convection film boiling heat, J. Heat Transfer ASME 111 (1989) 760–
600 °C. Special instrumentations and signal treatment were set 766.
M. Gradeck et al. / International Journal of Heat and Mass Transfer 54 (2011) 5527–5539 5539

[22] J.V. Filipovic, Film boiling over a moving horizontal surface, Ph.D. Purdue [35] Y. Mitsutake, M. Monde, Ultra high critical heat flux during forced flow boiling
University, 1994. heat transfer with an impinging jet, J. Heat Transfer 125 (6) (2003) 1038–1045.
[23] Z.-H. Liu, J. Wang, Study on film boiling heat transfer for water jet impinging [36] J.A. Hammad, Y. Mitsutake, M. Monde, Movement of maximum heat flux and
on high temperature flat plate, Int. J. Heat Mass Transfer 44 (2001) 2475– wetting front during quenching of hot cylindrical block, Int. J. Therm. Sci. 43
2481. (8) (2004) 743–752.
[24] Z-H. Liu, Prediction of minimum heat flux for water jet boiling on a hot plate, J. [37] D.H. Wolf, F.P. Incropera, R. Viskanta, Local jet impingement boiling heat
Thermophys. Heat Transfer 17 (2003) 2. transfer, Int. J. Heat Transfer 39 (7) (1996) 1395–1406.
[25] H. Robidou, Etude expérimentale du refroidissement diphasique à haute [38] J. Filipovic, R. Viskanta, F.P. Incropera, Cooling of a moving steel strip by an
température par jet d’eau impactant, Thèse de doctorat, Université Henri array of round jets, in: 35th MWSP Conference Proceedings, ISS-AIME, vol.
Poincaré, Nancy I, 2000. XXXI, 1994, p. 317.
[26] N. Hatta, J-I. Kokado, H. Takuda, J. Harada, K. Hiraku, Predictable modeling for [39] J. Filipovic, R. Viskanta, F.P. Incropera, T.A. Veslocki, Cooling of a moving steel
coolings process of a hot steel plate by a laminar water bar, Arch. strip by an array of round jets, Steel Res. (12) (1994).
Eisenhüttenwes. 55 (1984). Nr.4. [40] F. Volle, D. Maillet, M. Gradeck, A. Kouachi, M. Lebouché, Practical application
[27] J.-I. Kokado, N. Hatta, H. Takuda, J. Harada, N. Yasuhira, An analysis of film of inverse heat conduction for wall condition estimation on a rotating cylinder,
boiling phenomena of subcooled water spreading radially on hot steel plate, Int. J. Heat Mass Transfer 52 (1-2) (2009) 210–221. 15.
Arch. Eisenhüttenwes. 55 (3) (1984). März. [41] A. Kouachi, Étude expérimentale de l’ébullition convective d’un jet d’eau plan
[28] N. Hatta, H. Osakabe, Numerical modeling for cooling process of a moving hot impactant une surface mobile portée à hautes températures, Thèse de
plate by a laminar water curtain, ISIJ Int. 29 (11) (1995) 919–925. Doctorat de l’Université de Nancy 1, Nancy, 2006. ddc: 620. <http://
[29] N. Karwa, T. Gambaryan-Roisman, P. Stephan, C. Tropea, A hydrodynamic www.scd.uhp-nancy.fr/docnum/SCD_T_2006_0136>.
model for subcooled liquid jet impingement at the Leidenfrost condition, Int. J. [42] D. Maillet, S. André, J.C. Batsale, A. Degiovanni, C. Moyne, Thermal
Therm. Sci. (2011), doi:10.1016/j.ijthermalsci.2011.01.021. Quadrupoles: Solving the Heat Equation through Integral Transforms, John
[30] P.J. Berenson, Film boiling heat transfer from a horizontal surface, J. Heat Wiley and Sons, UK, 2000.
Transfer 83 (3) (1961) 351–358. [43] F. Volle, D. Maillet, M. Gradeck, M. Lebouché, Semi-analytical inverse heat
[31] E.K. Kalinin, I.I. Berlin, V.V. Kostiouk, Transition boiling heat transfer, Advances conduction on a rotating cylinder with Laplace and Fourier transforms, Inverse
in Heat Transfer, vol. 18, Academic Press, 1987. Problem in Science and Engineering-IPSE-Cambridge 16 (5) (2008) 655–674.
[32] Pan Chin, J.Y. Hwang, T.L. Lin, The mechanism of heat transfer in transition [44] F. Volle, M. Gradeck, A. Kouachi, D. Maillet, M. Lebouché, Inverse heat
boiling, Int. J. Heat Mass Transfer 32 (7) (1989) 1337–1349. conduction applied to the measurement of heat fluxes on a rotating cylinder:
[33] N. Nagai, S. Nishio, Leidenfrost temperature on an extremely smooth surface, comparison between an analytical and a numerical technique, J. Heat Transfer
Exp. Therm. Fluid Sci. 12 (1996) 373–379. 130 (8) (2008), doi:10.1115/1.2928013.51-52.081302.
[34] H. Ohtake, Y. Koizumi, Derivations of correlation and liquid–solid contact [45] N. Seiler-Marie, J.M. Seiler, O. Simonin, Transition boiling at jet impingement,
model of transition boiling heat transfer, JSME Int. J. Ser. B (2) (2006). Int. J. Heat Mass Transfer 47 (2004) 5059–5070.

You might also like